Airfoil Thickness Effects on the Thrust Generation of Plunging Airfoils u,

advertisement
JOURNAL OF AIRCRAFT
Vol. 49, No. 5, September–October 2012
Airfoil Thickness Effects on the Thrust Generation
of Plunging Airfoils
Meilin Yu,∗ Z. J. Wang,† and Hui Hu‡
Iowa State University, Ames, Iowa 50011
Downloaded by IOWA STATE UNIVERSITY on February 2, 2013 | http://arc.aiaa.org | DOI: 10.2514/1.C031720
DOI: 10.2514/1.C031720
A numerical study was conducted to investigate the effects of airfoil thickness on the thrust generation of plunging
airfoils and to assess the contributions of pressure and viscous forces in flapping propulsion. A series of NACA
symmetric airfoils with thickness ratio ranging from 4.0% to 20.0% of the airfoil chord length were used in the
present study to undertake a same sinusoid plunging motion at a low Reynolds number of Re 1200 with the
plunging Strouhal number Str 0:45 and reduced frequency k 3:5. It was found that the thickness of the airfoils
would affect the evolution of the unsteady vortex structures around the plunging airfoils significantly, even though
the airfoils were set to undertake the same plunging motion. The different behaviors of the unsteady vortex structures
shedding from the airfoils with different thickness were found to cause dramatic changes to the resultant
aerodynamic force acting on the plunging airfoils. For a thick plunging airfoil with its thickness ratio greater than
9.0%, pressure force was found to play a dominant role in the thrust generation, and viscous force would be almost
negligible and contribute mainly to drag production. It confirms that the traditional inviscid model of the Knoller–
Betz effect (i.e., ignoring viscous effect) can be used to explain many phenomena associated with flapping propulsion.
A new finding of the present study is the substantial contribution of viscous force to the thrust generation for thin
plunging airfoils (i.e., the thickness ratio less than 8.0%). Viscous force was found to become thrust-producing,
instead of drag-producing, and it played a nonnegligible role in the thrust generation for the thin airfoils (i.e., viscous
force would produce up to 20.5% of the total thrust for NACA0004 airfoil in the present study). The role change of
viscous force in the thrust generation of the plunging airfoils was found to be closely related to the variations of the
dynamics of the unsteady vortex structures around the plunging airfoils.
flapping-wing-based MAVs. For example, while birds and insects
flap their thin wings to fly (i.e., wing thickness is only a few percent of
the chord length), much thicker airfoils (airfoil thickness >10% of
chord length) were usually used in previous studies to reveal the
underlying physics of flapping flight [6–12]. Although numerous
experimental and numerical studies have been conducted recently to
investigate the effects of kinematic parameters of flapping motions
(such as the flapping frequency, amplitude, and phase difference
between plunging and pitching motions) on the thrust generation and
propulsive efficiency, the influence of airfoil thickness on flapping
propulsion has not yet been fully explored [10]. Furthermore, while
the inviscid model of Knoller–Betz effect (i.e., ignoring viscous
effect) has been used widely to explain many phenomena associated
with flapping propulsion [8–10], the role of viscous force in flapping
propulsion is still poorly understood. Many fundamental questions
still remain to be answered, such as “Are viscous effects negligible
for flapping propulsion under all conditions?” and “Although
viscous force is known to be usually drag-producing, can it ever
make a positive contribution to the thrust generation in flapping
propulsion?”
In this short paper, we report a numerical study to investigate the
effects of airfoil thickness on the thrust generation of plunging
airfoils and to assess the contribution of viscous force to the thrust
generation in flapping propulsion. A series of commonly used
symmetric NACA airfoils with thickness ranging from 4 to 20% of
the chord length were used to undertake the same plunging motion at
a low Reynolds number of Re 1200. The behavior of the unsteady
vortex structures around the plunging airfoils and the resultant
aerodynamic forces acting on the plunging airfoils were compared
quantitatively to reveal the underlying physics related to flapping
propulsion. The contribution of viscous force on the thrust
generation in flapping propulsion was also examined in detail based
on the quantitative comparison.
Introduction
M
ICRO air vehicles (MAVs) have been one of the most active
research topics in the aerospace engineering community in
recent years. The miniaturized aircraft is expected to open up new
opportunities for surveillance-like missions, especially in hazardous
environments inaccessible to ground vehicles. Among different
MAV designs, flapping-wing-based designs stand out with high
efficiency and excellent maneuverability, as demonstrated by the
natural fliers such as birds and insects. It has long been realized that
steady-state aerodynamics does not accurately account for the
aerodynamic forces produced in flapping flight. This has prompted
extensive studies to elucidate the fundamental mechanism of
flapping flight to produce enough aerodynamic forces needed for
propulsion and maneuvering. Knoller [1] and Betz [2] are among the
first to propose an invisid theory, which is known as the Knoller–Betz
effect, to explain why a flapping wing can generate thrust in flapping
motion. Katzmayr [3] provided the first experimental verification
of the Knoller–Betz effect by placing a stationary wing into a
sinusoidally oscillating airflow. Ober [4] provided additional
theoretic explanations and calculations to confirm Katzmayr’s
experimental results. Much progress has been made since then to
uncover the underlying physics of flapping propulsion [5–15].
Although many important findings have been derived through
those previous studies, much work is still needed for a better
understanding of flapping propulsion for the optimum design of
Received 26 October 2011; revision received 21 February 2012; accepted
for publication 22 February 2012. Copyright © 2012 by Meilin Yu, Z. J.
Wang, and Hui Hu. Published by the American Institute of Aeronautics and
Astronautics, Inc., with permission. Copies of this paper may be made for
personal or internal use, on condition that the copier pay the $10.00 per-copy
fee to the Copyright Clearance Center, Inc., 222 Rosewood Drive, Danvers,
MA 01923; include the code 0021-8669/12 and $10.00 in correspondence
with the CCC.
∗
Graduate Student, Department of Aerospace Engineering.
†
Professor, Department of Aerospace Engineering. Associate Fellow
AIAA.
‡
Associate Professor, Department of Aerospace Engineering; huhui@
iastate.edu. Associate Fellow AIAA.
Numerical Method and Studied Parameters
In the present study, a high-order spectral difference method with
dynamic unstructured grids was used for the numerical simulation.
1434
1435
YU, WANG, AND HU
The governing equations for the fluid flow are the unsteady
Navier–Stokes equations in a conservation form, which can be
expressed as
Downloaded by IOWA STATE UNIVERSITY on February 2, 2013 | http://arc.aiaa.org | DOI: 10.2514/1.C031720
@Q @F @G @H
0
@x @y
@z
@t
(1)
Herein, Q ; u; v; w; ET are the conservative variables; is the fluid density; u, v, and w are the Cartesian velocity
components; and E is the total initial energy. F, G, H are the total
fluxes including both the inviscid and viscous flux vectors,
i.e.,F Fi Fv , G Gi Gv and H Hi H v . Detailed
formulas for the fluxes can be found in Yu et al. [15]. With
the assumption that the fluid obeys the perfect gas law, the
pressure is related to the total initial energy by E p=
1 12u2 v2 w2 , which closes the solution system.
To achieve an efficient implementation, a time-dependent
coordinate transformation from the physical domain t; x; y; z to the
computational domain ; ; ; is applied on Eq. (1), which is
@Q~ @F~ @G~ @H~ 0
@x @y
@z
@
(2)
where
8
Q~ jJjQ
>
>
>
>
< F~ jJjQ Fx Gy Hz >
> G~ jJjQ Fx Gy Hz >
>
: ~
H jJjQ Fx Gy Hz (3)
Herein, t and ; ; 2 1; 13 are the local coordinates in
the computational domain. In the transformation shown previously,
the Jacobian matrix J takes the following form:
0
1
x x x z
@x; y; z; t B
y y y y C
C
(4)
B J
@; ; ; @ z z z z A
0 0 0 1
*
Note that the grid velocity vg xt ; yt ; zt is related with
; ; by
8
*
>
>
< vg r
*
(5)
vg r
>
>
*
:
vg r
In the present study, H-refinement (grid refinement) and prefinement studies were conducted at first to determine the suitable
grid and numerical accuracy. Based on the investigations, a thirdorder-accurate scheme with a medium mesh was chosen. A timerefinement study was also performed to determine a reasonable
nondimensional time step for the present study. Further information
about the implementation of the method described previously for the
numerical simulation of the unsteady flows around flapping airfoils
as well as the validation of the simulation results against experimental data is available in [15].
The airfoils used in the present study are a series of symmetric
NACA airfoils, i.e., NACA0004, NACA0006, NACA0009,
NACA0012, and NACA0020 airfoils. The airfoils were set to
undertake a plunging motion, which can be expressed as y
A sin2ft, where f is the flapping frequency, and A is the plunging
amplitude. The Reynolds number Re, based on the airfoil chord
length C and the freestream velocity V1, was set to be 1200 for the
present study, i.e., Re V1 C= 1200, which is well within the
insect flight regime. Strouhal number Str 2fA=V1 and reduced
frequency k 2fC=V1 are the most commonly used nondimensional parameters to characterize the kinematics of flapping airfoils/
wings. In the present study, the Strouhal number of the plunging
airfoils was chosen to be 0.45, i.e., Str 0:45. The reduced
frequency of the plunging motion was set to be 3.5, i.e., k 3:5. It
has been suggested that the wake flow downstream of a flapping
airfoil/wing can be characterized as drag-producing, neutral, or
thrust-producing, depending on the flapping frequency and stroke
amplitude [6–14]. Based on the findings of the previous work of
Jones et al. [8] and Lewin and Haj-Hariri [12], with the kinematic
parameters used in the present study, the wake flows downstream the
plunging airfoils should be thrust-producing, which was confirmed
by the numerical simulation results of the present study.
Results and Discussions
Figures 1 and 2 display the typical behaviors of the unsteady
vortex structures around a thick airfoil (e.g., NACA0020) and a thin
airfoil (e.g., NACA0004) in a plunging cycle. The flow pattern
around the plunging airfoils at such a relatively large Strouhal
number (i.e., Str 0:45) was found to be featured mainly by the
periodic shedding of leading-edge vortices (LEVs) and trailing-edge
vortices (TEVs) as well as the interactions among LEVs, TEVs, and
plunging airfoils, which agrees with those reported by Lewin and
Haj-Hariri [12]. It is also observed that, even though the airfoils were
set to undertake the same plunging motion, the evolutions of the
unsteady LEVs and TEVs around the plunging airfoils were found to
vary significantly due to the thickness differences of the airfoils. As
shown in Fig. 1, for the thick airfoil case, the LEVs shed from the
airfoil leading edge were found to travel downstream along with the
freestream continuously and then interact with the TEVs further
downstream. Similar behavior of the LEVs was also reported by
Ashraf et al. [10] in their study of the vortex structures around a
plunging NACA0012 airfoil. However, for the thin airfoil (e.g.,
NACA0004) as shown in Fig. 2, instead of traveling downstream
along with the freestream continuously, the LEVs over the lower (or
upper) surface of the airfoil were found to stay close to the airfoil
leading edge during the entire downstrokes (or upstrokes) of the
plunging motion. After being stretched seriously, the LEVs were
found to move against the freestream around the sharp airfoil leading
edge and shift to the upper (or lower) side of the airfoil during the
subsequent upstrokes (or downstrokes). Such phenomena were also
found by Lewin and Haj-Hariri [12] and were named as LEV
circumnavigation. Associated with the different behaviors of the
LEVs around thick and thin airfoils, the flow patterns and the
resultant aerodynamic forces acting on the plunging airfoils were
also found to vary dramatically.
Figure 3 shows the pressure distributions and velocity vector fields
(only 2% of the vectors were shown) around NACA0020 and
NACA0004 airfoils at the same plunging phase angle of 180 deg. It can be found that, for a thick airfoil such as NACA0020,
corresponding to the rolling up of the LEVs on the airfoil upper (or
lower) surface during the downstrokes (or upstrokes) of the plunging
motion, a region with relatively low pressure was found on the airfoil
upper (or lower) surface near the airfoil leading edge as shown in
Fig. 3a, which is favorable for the thrust generation. The region with
relatively low pressure was found to separate from the airfoil surface
and move downstream as the LEVs shed from the airfoil leading edge
and travel downstream. However, for a thin airfoil as shown in
Fig. 3b, associated with the LEV circumnavigation described
previously, low-pressure regions were found to exist on both the
upper and lower surfaces, and the low-pressure regions would remain
near the airfoil leading edge during almost the whole plunging cycle.
Furthermore, the LEV circumnavigation was also found to induce
strong reversed flows on both the upper and lower surfaces near the
leading edge of the thin airfoils. Corresponding to the strong reversed
flows near the airfoil leading edge, the viscous force acting on the
upper and lower surfaces of the thin airfoil could actually be thrustproducing, instead of being drag-producing. The interesting finding
can be revealed more quantitatively in the analysis of the resultant
aerodynamic forces acting on the plunging airfoils.
Based on the distributions of pressure and viscous forces on the
surfaces of the plunging airfoils, the resultant aerodynamic forces
acting on the plunging airfoils in the term of thrust coefficient
2
S were determined. Figure 4 shows the
CT Thrust=0:5V1
histories of the thrust coefficients of the five plunging airfoils
Downloaded by IOWA STATE UNIVERSITY on February 2, 2013 | http://arc.aiaa.org | DOI: 10.2514/1.C031720
1436
YU, WANG, AND HU
Fig. 1
Evolution of the unsteady vortex structures around NACA0020 airfoil in a plunging cycle.
investigated in the present study. It can be seen that, although the
thrust coefficients of the plunging airfoils were found to fluctuate
greatly in each plunging cycles, the fluctuating amplitude of the
thrust coefficients decreases as the thickness of the plunging airfoil
decreases. Although all the airfoils were set to undertake the same
periodic plunging motion, only the thrust coefficients of the thicker
airfoils (e.g., NACA0020 and NACA0020 airfoils) were found to be
periodic, as expected. The thrust coefficients of the thinner airfoils
(e.g., NACA0004 and NACA0006 airfoils) were found to become
aperiodic even though the plunging motion of the airfoils is periodic.
The aperiodic behavior of the flowfield around a plunging airfoil was
also reported by Lewin and Haj-Hariri [12] with an elliptical airfoil
plunging at a similar Strouhal numbers (e.g., Str 0:48) as that of
the present study. It should also be noted that, for the thicker airfoils
(e.g., NACA0012 or NACA0020), while the resultant aerodynamic
force acting on the plunging airfoils were found to be thrustproducing for most of the time in each plunging cycles (i.e., thrust
coefficient being positive), the resultant aerodynamic force could
also become drag-producing (i.e., thrust coefficient becoming
negative) at some phase angles. However, the thrust coefficients of
the thinner airfoils (e.g., NACA0004 or NACA0006) were found to
be positive almost in the entire plunging cycles, which indicates that
almost no drag was experienced by the thinner airfoils during the
plunging motion. This is believed to be closely related to the LEV
circumnavigation to maintain low-pressure regions and reverse flows
near the leading edges of the thinner airfoils.
To assess the role of viscous force in the thrust generation of
flapping propulsion, the total thrusts acting on the plunging airfoils
were decomposed into
R two parts, i.e., one part contributed from
pressure
R force, Tp pnx ds, and the other from the viscous force,
denotes the viscous
Tv xx nx xy ny xz nz ds, where
stresses. Figure 5 shows the comparisons of the total thrust
coefficients (i.e., considering the contributions from both pressure
and viscous forces) and the thrust coefficients based on the
contribution from pressure force only (i.e., ignoring the viscous
forces) for NACA0020 and NACA0004 airfoils. Although the
profiles in solid lines represent the total thrust coefficients, the
profiles in dashed lines indicate the results based on the contribution
from pressure force only. The differences between the solid and
dashed lines would represent the contribution of viscous force on the
thrust generation. It can be seen clearly that, compared with viscous
force, pressure force was found to play a dominant role in the thrust
generation of flapping propulsion. It is also observed that the
differences between the solid and dashed lines were found to become
larger for the thinner airfoil (e.g., NACA0004), compared with those
for the thicker airfoil case (e.g., NACA0020). This indicates that the
effects of viscous force on the thrust generation in flapping
propulsion would become stronger for the thinner airfoils.
To reveal the contribution of viscous force on the thrust generation
of the plunging airfoils more clearly and quantitatively, the averaged
total thrust coefficients of the plunging airfoils, hCT i, and the
contributions from pressure force, hCTP i, and viscous force, hCTV i,
over plunging cycles were calculated. As revealed clearly from the
results listed in Table 1, the averaged thrust coefficients of the
plunging airfoils vary significantly as the airfoil thickness changes.
With the plunging kinematic parameters and the airfoil thickness
Downloaded by IOWA STATE UNIVERSITY on February 2, 2013 | http://arc.aiaa.org | DOI: 10.2514/1.C031720
YU, WANG, AND HU
Fig. 2
1437
Evolution of the unsteady vortex structures around NACA0004 airfoil in a plunging cycle.
range used in the present study, a thicker airfoil was found to generate
a larger averaged thrust when undertaking a same plunging motion.
The finding was found to agree with the conclusion reported by
Ashraf et al. [10]. The results shown in Table 1 also confirmed that
pressure force would play a dominant role in the thrust generation of
flapping propulsion. For the thicker airfoils (e.g., NACA0012 and
NAC0020 airfoils), viscous force was found to be mainly dragproducing, and its effect was found to be very small ( 2:0%), which
Fig. 3 Comparison of the pressure distributions and velocity fields around NACA0020 and NACA0004 airfoils at the phase angle of 180 deg in the
plunging motion.
1438
YU, WANG, AND HU
Downloaded by IOWA STATE UNIVERSITY on February 2, 2013 | http://arc.aiaa.org | DOI: 10.2514/1.C031720
Conclusions
Fig. 4 Histories of the thrust coefficients of the NACA symmetrical
airfoils in plunging motion.
Fig. 5 Comparison of the total thrust coefficients of the plunging
airfoils and the contributions from pressure force only.
is negligible. The finding can be used to explain why the traditional
inviscid model of Knoller–Betz effect (i.e., ignoring the viscous
effect) can be used to explain many phenomena associated with
flapping propulsion. The drag-producing nature of viscous force for
thick airfoils was found to agree with the results reported by Wang
[11]. More interestingly, as revealed from the results given in
Table 1, viscous force becomes thrust-producing, instead of
drag-producing, for the thinner airfoils (e.g., NACA0006 and
NACA0004). The contribution of viscous force to the thrust
generation in flapping flight was found to become more and more
substantial as the thickness of the plunging airfoil decreases (i.e.,
13.6% for NACA0006 and up to 20.5% for NACA0004). This is
believed to be closely related to the existence of reverse flows near the
leading edges on both sides of the airfoil surfaces during almost
whole plunging cycles for the thinner airfoils, as shown clearly in
Fig. 3.
Table 1 Averaged total thrust coefficients hCT i
and the contributions from the pressure
force hCTP i and viscous force hCTv i
Airfoil
hCT i
hCTP i
hCTV i
hCTV i=hCT i
NACA0004
NACA0006
NACA0009
NACA0012
NACA0020
0.261
0.384
0.410
0.573
0.920
0.208
0.332
0.421
0.584
0.937
0.053
0.052
0:011
0:011
0:017
20.5%
13.6%
2:6%
1:9%
1:9%
A numerical study was conducted to investigate the effects of
airfoil thickness on the thrust generation of plunging airfoils and to
assess the contribution of viscous force to flapping propulsion. A
series of commonly-used NACA symmetric airfoils were used in the
present study to undertake a same plunging motion at a low Reynolds
number of Re 1200 with the plunging Strouhal number Str 0:45 and reduced frequency k 3:5. It was found that, even though
the airfoils were set to undertake the same plunging motion, the
evolutions of the vortex structures around the plunging airfoils and
the resultant aerodynamic forces acting on the airfoils varied
dramatically due to the difference in airfoil thickness.
Although the leading-edge vortices (LEVs) of the thicker airfoils
(e.g., >9:0% thickness ratio for the present study) were found to shed
periodically and travel downstream along with the freestream
continuously, LEV circumnavigation was found for the thinner
airfoils (e.g., <9:0% thickness ratio) with LEVs stretched and
remaining near the airfoil leading edges for most of time in the
plunging cycles. Associated with the LEV circumnavigation, lowpressure regions and reverse flows were found to remain near the
leading edge on both sides of the airfoils during almost whole
plunging cycles for the thinner airfoils. The different behaviors of the
LEVs for the thick and thin airfoils were found to affect the resultant
aerodynamic force acting on the plunging airfoils dramatically.
Although the thrust generation of the plunging airfoils with larger
thickness was found to be periodic as expected, the thrust generation
of the thin airfoils was found to become aperiodic even though the
plunging motion is periodic.
The present study also revealed that pressure force plays a
dominant role in the thrust generation of the plunging airfoils. As
expected, viscous force contributes mainly to drag production, and
its effect is almost negligible for the airfoils with relatively large
thickness, which explains why the traditional inviscid model of
Knoller–Betz effect (i.e., ignoring the viscous effect) can be used to
explain many phenomena associated with flapping propulsion.
Another important finding of the present study is the substantial
contribution of viscous force to the thrust generation for thin airfoils.
The viscous force was found to become thrust-producing, instead of
drag-producing, and it played a nonnegligible role in the thrust
generation for the thin airfoils (i.e., producing up to 20.5% of the total
thrust for NACA0004 airfoil in the present study). The role change of
viscous force in the thrust generation of the plunging airfoils was
found to be closely related to the dynamics of the unsteady vortex
structures around the plunging airfoils.
Acknowledgments
The authors want to thank Feng Liu of the University of California,
Irvine, and JC Wu and Hong Liu of Shanghai Jiao Tong University
for helpful discussions related to the present study. The support of the
National Science Foundation under award number CBET-1064235
is gratefully acknowledged.
References
[1] Knoller, R., “Die Gesetze des Luftwiderstandes,” Flug- und
Motortechnik, Vol. 3, No. 21, 1909, pp. 1–7.
[2] Betz, A., “Ein Beitrag zur Erklaerung des Segelfluges,” Zeitschrift fuer
Flugtechnik und Motorluftschiffahrt, Vol. 3, 1912, pp. 269–270.
[3] Katzmayr, R., “Effect of Periodic Changes of Angle of Attack on
Behavior of Airfoils,” NACA TM 147, 1922.
[4] Ober, S., “Note on the Kaztmayr Effect on Airfoil Drag,” NACA
TN 214, 1925.
[5] von Kármán, T., and Burgers, J. M., Aerodynamic Theory: A General
Review of Progress, edited by W. F. Durand, Vol. 2, Springer, Berlin,
1935.
[6] Koochesfahani, M. M., “Vortical Patterns in the Wake of an Oscillating
Airfoil,” AIAA Journal, Vol. 27, 1989, pp. 1200–1205.
doi:10.2514/3.10246
[7] Dickinson, M. H., Lehmann, F. O., and Sane, S. P., “Wing Rotation and
the Aerodynamic Basis of Insect Flight,” Science, Vol. 284, 1999,
pp. 1954–1960.
doi:10.1126/science.284.5422.1954
YU, WANG, AND HU
Downloaded by IOWA STATE UNIVERSITY on February 2, 2013 | http://arc.aiaa.org | DOI: 10.2514/1.C031720
[8] Jones, K. D., Dohring, C. M., and Platzer, M. F., “Experimental and
Computational Investigation of the Knoller–Betz Effect,” AIAA
Journal, Vol. 36, No. 7, 1998, pp. 1240–1246.
doi:10.2514/2.505
[9] Platzer, M. F., Jones, K. D., Young, J., and Lai, J. C. S., “Flapping-Wing
Aerodynamics: Progress and Challenges,” AIAA Journal, Vol. 46,
No. 9, 2008, pp. 2136–2149.
doi:10.2514/1.29263
[10] Ashraf, M. A., Yong, J., and Lai, J. C. S., “Reynolds Number, Thickness
and Camber Effects on Flapping Airfoil Propulsion,” Journal of Fluids
and Structures, Vol. 27, 2011, pp. 145–160.
doi:10.1016/j.jfluidstructs.2010.11.010
[11] Wang, Z. J., “Vortex Shedding and Frequency Selection in Flapping
Flight,” Journal of Fluid Mechanics, Vol. 410, 2000, pp. 323–341.
doi:10.1017/S0022112099008071
1439
[12] Lewin, G. C., and Haj-Hariri, H., “Modeling Thrust Generation of a
Two-Dimensional Heaving Airfoil in a Viscous Flow,” Journal of Fluid
Mechanics, Vol. 492, 2003, pp. 339–362.
doi:10.1017/S0022112003005743
[13] Shyy, W., Lian, Y., Tang, J., Viieru, D., and Liu, H., Aerodynamics of
Low Reynolds Number Flyers, Cambridge Univ. Press, New York,
2008.
[14] Hu, H., Clemons, L., and Igarashi, H., “An Experimental Study of the
Unsteady Vortex Structures in the Wake of a Root-Fixed Flapping
Wing,” Experiments in Fluids, Vol. 51, No. 2, 2011, pp. 347–359.
doi:10.1007/s00348-011-1052-z
[15] Yu, M. L., Wang, Z. J., and Hu, H., “A High-Order Spectral Difference
Method for Unstructured Dynamic Grids,” Computers and Fluids,
Vol. 48, 2011, pp. 84–97.
doi:10.1016/j.compfluid.2011.03.015
Download